Download The Hydroxypyruvate-Reducing System in Arabidopsis: Multiple

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

History of genetic engineering wikipedia , lookup

Transcript
The Hydroxypyruvate-Reducing System in Arabidopsis:
Multiple Enzymes for the Same End1[W]
Stefan Timm*, Alexandra Florian, Kathrin Jahnke, Adriano Nunes-Nesi2,
Alisdair R. Fernie, and Hermann Bauwe
Department of Plant Physiology, University of Rostock, D–18059 Rostock, Germany (S.T., K.J., H.B.); and Max
Planck Institute of Molecular Plant Physiology, D–14476 Golm, Germany (A.F., A.N.-N., A.R.F.)
Hydroxypyruvate (HP) is an intermediate of the photorespiratory pathway that originates in the oxygenase activity of the key
enzyme of photosynthetic CO2 assimilation, Rubisco. In course of this high-throughput pathway, a peroxisomal transamination
reaction converts serine to HP, most of which is subsequently reduced to glycerate by the NADH-dependent peroxisomal
enzyme HP reductase (HPR1). In addition, a NADPH-dependent cytosolic HPR2 provides an efficient extraperoxisomal
bypass. The combined deletion of these two enzymes, however, does not result in a fully lethal photorespiratory phenotype,
indicating even more redundancy in the photorespiratory HP-into-glycerate conversion. Here, we report on a third enzyme,
HPR3 (At1g12550), in Arabidopsis (Arabidopsis thaliana), which also reduces HP to glycerate and shows even more activity with
glyoxylate, a more upstream intermediate of the photorespiratory cycle. The deletion of HPR3 by T-DNA insertion
mutagenesis results in slightly altered leaf concentrations of the photorespiratory intermediates HP, glycerate, and glycine,
indicating a disrupted photorespiratory flux, but not in visible alteration of the phenotype. On the other hand, the combined
deletion of HPR1, HPR2, and HPR3 causes increased growth retardation, decreased photochemical efficiency, and reduced
oxygen-dependent gas exchange in comparison with the hpr1xhpr2 double mutant. Since in silico analysis and proteomic
studies from other groups indicate targeting of HPR3 to the chloroplast, this enzyme could provide a compensatory bypass for
the reduction of HP and glyoxylate within this compartment.
Photorespiration is a high-throughput pathway that
is essential for the normal growth and development of
plants in air (for review, see Bauwe et al., 2010). This
process is caused by Rubisco, which cannot fully
discriminate between carbon dioxide and oxygen.
Carboxylation of ribulose 1,5-bisphosphate by Rubisco
yields two molecules, 3-phosphoglycerate (3PGA),
which can directly enter the Calvin cycle. In contrast,
the oxygenase activity of Rubisco produces one molecule each of 3PGA and 2-phosphoglycolate (2PG).
The latter compound inhibits several essential enzymes, such as triosephosphate isomerase and phosphofructokinase (Anderson, 1971; Kelly and Latzko,
1976), and is toxic for plant cells. The photorespiratory
pathway recovers and detoxifies 2PG by converting it
back into the Calvin cycle intermediate, 3PGA. This
requires the interplay of several enzymes located in
1
This work was supported by the Landesgraduiertenförderung
Mecklenburg-Vorpommern (to S.T.) and by the Deutsche Forschungsgemeinschaft (grant nos. BA 1177/7 and BA 1177/12 within
the research unit FOR 1186 PROMICS to H.B.).
2
Present address: Departamento de Biologia Vegetal, Universidade Federal de Viçosa, 36570–000 Viçosa, Minas Gerais, Brazil.
* Corresponding author; e-mail [email protected].
The author responsible for distribution of materials integral to the
findings presented in this article in accordance with the policy
described in the Instructions for Authors (www.plantphysiol.org) is:
Hermann Bauwe ([email protected]).
[W]
The online version of this article contains Web-only data.
www.plantphysiol.org/cgi/doi/10.1104/pp.110.166538
694
the chloroplast, the peroxisome, the mitochondrion,
and the cytosol.
One of the key intermediates of the photorespiratory
cycle is hydroxypyruvate (HP), which is produced in
the peroxisome from Ser by the enzyme Ala:glyoxylate
(Glx) aminotransferase (AGT1; Liepman and Olsen,
2001). HP then is mostly reduced by the peroxisomal NADH-dependent HP reductase (HPR1) to
D-glycerate and to a lesser extent by a cytosolic NADPHdependent HPR2 (Timm et al., 2008). The resulting
glycerate is phosphorylated to 3PGA by the chloroplastidial enzyme glycerate 3-kinase (Boldt et al., 2005)
and reenters the Calvin cycle.
Deletion of photorespiratory enzymes typically
leads to a strong air sensitivity of the respective mutants, which, however, can be fully recovered by
elevated-CO2 conditions. While this is a distinctive
feature of most photorespiratory mutants (Igarashi
et al., 2003; Boldt et al., 2005; Voll et al., 2006; Schwarte
and Bauwe, 2007), Arabidopsis (Arabidopsis thaliana)
HPR1 knockout mutants grow nearly normally in
ambient air with moderate photoperiods and show
only minor changes in photosynthetic and metabolic
parameters under these conditions. The additional
deletion of the cytosolic HPR2 distinctly elevates the
oxygen sensitivity, but this hpr1xhpr2 double mutant
can still survive long-term exposure to ambient air
(Timm et al., 2008). Since this indicates even more
redundancy in the photorespiratory HP-into-glycerate
conversion step, although with a low apparent capacity, we searched the Arabidopsis genome (Arabidopsis
Plant PhysiologyÒ, February 2011, Vol. 155, pp. 694–705, www.plantphysiol.org Ó 2010 American Society of Plant Biologists
Downloaded from on August 3, 2017 - Published by www.plantphysiol.org
Copyright © 2011 American Society of Plant Biologists. All rights reserved.
Photorespiratory Hydroxypyruvate Metabolism in Arabidopsis
Genome Initiative, 2000) for related enzymes and identified protein At1g12550 as another putative HPR.
Studies with the recombinant protein, which shows up
to approximately 50% sequence identity with HPR1
and HPR2 and is annotated as a putative D-isomerspecific 2-hydroxyacid dehydrogenase, confirmed that
At1g12550 is able to reduce HP and Glx in vitro. A
physiological characterization of knockout mutants of
this protein revealed minor but distinct changes in the
photorespiratory cycle, as evidenced by slightly increased levels of pathway intermediates. This notwithstanding, the fact that the knockout had no effect on
growth in normal air indicates that in the presence of
HPR1 and HPR2, this protein plays only a minor role
in the photorespiratory process. The combined deletion of HPR1, HPR2, and the newly identified HPRlike enzyme (hereafter referred to as HPR3) leads to
strong impairment of growth in normal air and a
massive air sensitivity after transfer from nonphotorespiratory conditions to air. Moreover, gas-exchange
parameters are strongly affected in the isolated triple
mutant, and we found further increased levels of Ser
and related metabolites of this photorespiratory amino
acid. Since HPR3 has been reported to be located in
chloroplasts both by the in silico prediction database
Aramemnon (Schwacke et al., 2003) and proteomic
studies (Yu et al., 2008), our results provide evidence
that HPR3 works as an HPR in plants and likely
represents an additional bypass both to known HPRs
and to glyoxylate reductases in chloroplasts.
RESULTS
Identification of a D-Isomer-Specific 2-Hydroxyacid
Dehydrogenase as an Alternative HPR
The nonlethal phenotype of the hpr1xhpr2 double
mutant in ambient air indicated that additional HPreducing enzymes could be present in Arabidopsis.
For this reason, we performed a BLAST search to
identify potential candidates for such enzymes. These
analyses led us to a protein that is encoded by the gene
At1g12550 and shares 45% sequence identity with the
cytosolic HPR2. This protein belongs to the D-isomerspecific 2-hydroxyacid dehydrogenase family and
contains a NADH/NADPH-binding site. To provide
biochemical evidence, we overexpressed the corresponding open reading frame in Escherichia coli and
tested different substrate/cosubstrate combinations.
The purified enzyme showed activity with both HP
and Glx, and NADPH was the favored cosubstrate
(Fig. 1). Moreover, oxalate inhibits the recombinant
enzyme up to 70% for each substrate/cosubstrate
combination. This inhibition was reported to be a characteristic for HPR2 but not for HPR1 (Kleczkowski
et al., 1991; Timm et al., 2008). Concerning these findings, we conclude that the identified protein is not a
typical HPR such as HPR1 but rather able to function
similarly to HPR2; therefore, we termed it HPR3.
Another candidate protein, At2g45630 (Timm et al.,
2008), did not show activity with HP or Glx and is
likely unrelated to photorespiration.
Isolation of T-DNA Insertional Mutants of AtHPR3
To examine the physiological function of HPR3, we
set out to isolate two independent T-DNA insertion
lines for the gene At1g12550. In these lines, hereafter
designated as hpr3-1 and hpr3-2, insertions are located
in the second exon and the first exon (Fig. 2A), respectively. First, homozygous lines were isolated by genomic PCR, and sequencing of the resulting fragments
was obtained from both lines. The complete knockouts
then were verified by reverse transcription (RT)-PCR
analysis, using primer pairs that span the T-DNA
insertion sites of the two loci. The 40S ribosomal protein
S16 mRNA (At2g09990) was used as an internal control
for the integrity of the generated cDNA. The mature
transcript was found to be present in wild-type plants
but was absent in both insertion mutants (Fig. 2B).
These results confirmed the complete knockout of the
At1g12550 gene in both T-DNA insertion lines.
Knockout of AtHPR3 Does Not Impair Growth But Leads
to Slightly Increased Levels of
Photorespiratory Intermediates
These HPR3-deficient mutant lines were grown on
soil under different photoregimes to allow a comprehensive phenotypic analysis. However, no phenotypic
Figure 1. Activity of recombinant
At1g12550 (HPR3) and its inhibition
by oxalate. A, Specific activity with HP
and Glx using cofactors NADH and
NADPH, respectively. B, Inhibition of
the recombinant enzyme by oxalate (2
mM). Shown are mean activities 6 SD
from three independent measurements
(mmol min21 mg21 protein). Asterisks
show significant differences according
to Student’s t test (P , 0.01).
Plant Physiol. Vol. 155, 2011
695
Downloaded from on August 3, 2017 - Published by www.plantphysiol.org
Copyright © 2011 American Society of Plant Biologists. All rights reserved.
Timm et al.
Figure 2. Isolation of HPR3 knockout mutants. A, Structure of the
HPR3 gene and positions of the T-DNA insertions of two independent
lines (gray bars represent the exons). B, HPR3-specific transcripts were
undetectable in both homozygous lines but were found to be present in
wild-type plants. Whole-leaf RNA was isolated and analyzed by RTPCR using S16 mRNA as a control for the integrity of generated cDNA.
This result was obtained for six plants (each line) and three technical
replicates.
produced from the triple heterozygous intermediate,
and plants again were grown under elevated-CO2
conditions (1% CO2 in air). This procedure was chosen
to ensure that all homozygous combinations could be
isolated, considering the conditional lethal phenotype
of most other photorespiratory mutants (Boldt et al.,
2005; Voll et al., 2006; Schwarte and Bauwe, 2007).
Finally, a triple homozygous mutant of HPR1, HPR2,
and HPR3 could by isolated under these conditions.
The transcription of all three HPRs was analyzed by
RT-PCR in order to verify the complete knockout of all
HPRs. In the corresponding triple mutant, no transcripts were detectable from all three different genes,
using S16 mRNA as an internal control for the integrity of the generated cDNA (Fig. 4).
As described above, during the crossing process of
hpr1-1, hpr2-1, and hpr3-1, all possible combinations of
double mutants also could be isolated. As visible in
Figure 5B, the double mutant hpr1xhpr3 does not show
any further growth retardation relative to the corresponding hpr1-1 single mutant, nor do the hpr2-1 and
differences were observed compared with the wildtype control, regardless of which photoperiod was
used (10/14-h, 12/12-h, or 16/8-h day/night cycle). To
examine the effect of knocking out HPR3 at the biochemical level, we next carried out metabolic analyses
of the respective mutants. For this purpose, plants
were grown on soil with a standard photoperiod of
10 h. After reaching growth stadium 5.1 (Boyes et al.,
2001), samples were taken in the middle of the light
period and analyzed by gas chromatography-mass
spectrometry according to Lisec et al. (2006). Despite
the wide range of unaltered metabolites (Supplemental Table S1), we found some slightly increased intermediates of the photorespiratory cycle, namely HP,
3PGA, glycerate, glycolate, and Gly (Fig. 3). We also
found that these changes were accompanied by correspondingly altered fluxes, as assessed by the redistribution of label following feeding of leaf discs
with [U-13C]Glc (Supplemental Table S2; SienkiewiczPorzucek et al., 2008). These alterations provide evidence
that the photorespiratory flux is slightly disrupted and
thus that the function of HPR3 is associated with
photorespiration.
Generation of a Triple HPR Mutant Leads to
Increased Air Sensitivity But Fully Recovers in
High-CO2 Conditions
Given that the double knockout of HPR1 and HPR2
is still viable in air, multiple mutants were generated
by integrating the newly identified HPR3 mutant by
conventional crossing. For this purpose, the isolated
line hpr3-1 was crossed with the previously isolated
hpr1xhpr2 double knockout (Timm et al., 2008) to
generate an HPR triple mutant and additionally to
gain all other possible combinations of HPR double
mutants. After crossing, the resulting T1 plants were
characterized by genomic PCR. Thereafter, seeds were
Figure 3. Alteration of the leaf content of selected metabolites in the
individual hpr3-1 and hpr3-2 knockout plants compared with the wild
type (ecotype Columbia [Col-0]). Plants were grown in normal air with
a 10/14-h day/night cycle. Leaves of six individual plants per line were
harvested in the middle of the photoperiod at developmental stage 5.1
(Boyes et al., 2001). Mutant-to-wild-type ratios 6 SD of relative mean
metabolite contents are shown with the mean wild-type values arbitrarily set to 1 (wild type, white bars; hpr3-1, light gray bars; hpr3-2,
dark gray bars). Asterisks show significant changes according to Student’s t test (P , 0.05).
696
Plant Physiol. Vol. 155, 2011
Downloaded from on August 3, 2017 - Published by www.plantphysiol.org
Copyright © 2011 American Society of Plant Biologists. All rights reserved.
Photorespiratory Hydroxypyruvate Metabolism in Arabidopsis
this enzyme is only responsible for a small proportion
of the HP-into-glycerate interconversion if HPR1 and
HPR2 are present. However, the absence of all three
HP-reducing enzymes further leads to strong impairment of growth under ambient air conditions compared with the double knockout of HPR1 and HPR2.
Increased levels of CO2 in air (more than 0.15%) fully
recover this oxygen-sensitive phenotype (Fig. 5C),
which clearly demonstrates that the phenotype of the
HPR triple mutant is related to photorespiration.
HPR Mutants Show Impaired Growth and Contain
Less Chlorophyll
Figure 4. Crossing of hpr1-1, hpr2-1, and hpr3-1 leads to functional
inactivation of the three HPR genes. RT-PCR analysis shows the
absence of the transcripts encoded by the three HPR genes in leaves
of the isolated triple mutant, whereas all three transcripts are present in
the wild-type control. S16 mRNA was used as an internal control for the
integrity of the produced cDNA. This result was obtained with three
independent plants per line and two replicate RT-PCR analyses each.
hpr3-1 single mutants and the hpr2xhpr3 double mutant show any retardation relative to the corresponding wild type. Only the hpr1xhpr2 double mutant
shows a clear growth reduction compared with the
hpr1-1 single mutant. As we can learn from the double
knockouts including HPR3 (hpr1xhpr3 and hpr2xhpr3),
To further substantiate the visible impairment of
growth, rosette diameters and chlorophyll contents
were determined from short-day-grown plants (10/
14-h day/night cycle). As visible in Figure 5 (A–C) and
Figure 6, growth reduction is exacerbated by the
increasing extent of HPR mutation. HPR1 mutants
show growth inhibition up to 46% compared with the
corresponding wild type under short-day conditions.
This effect is not present with photoperiods longer
than 12 h (Fig. 5E; Timm et al., 2008). The hpr1xhpr2
double knockout is inhibited up to 76% compared
with the wild-type control and furthermore is reduced
up to 57% compared with the hpr1 single mutant. The
hpr1xhpr2xhpr3 triple mutant generated here is even
Figure 5. The combined deletion of HPR1,
HPR2, and HPR3 leads to a significant decrease in plant growth but is recovered under
high-carbon conditions. Plants were grown on
soil in ambient air (except for C and D, right
panel, which used 1% CO2) with a day/night
cycle of 10/14 h (except for E, which used 12/
12 h). Photographs were taken after 8 weeks
of cultivation. A, Phenotypes of the HPR
single mutants (from left to right: hpr1-1,
hpr2-1, and hpr3-1) compared with the
wild-type control (Col-0). B, Phenotypes of
the generated HPR double mutants (from left
to right: hpr1-1xhpr2-1, hpr1-1xhpr3-1, and
hpr2-1xhpr3-1). C, Phenotype of the isolated
HPR triple mutant in normal air (left panel)
compared with HPR1 and the HPR1/HPR2
double mutant (middle panel) and under highcarbon conditions (1% CO2) compared with
the wild-type control (right panel). D, A stronger phenotype is visible under short days (10/
14 h; left panel) but could be compensated by
cultivation in elevated CO2 (1%; right panel).
E, This effect vanishes by cultivation with
longer photoperiods (more than 12 h of light).
HPR2 and HPR3 single mutants do not display
this effect.
Plant Physiol. Vol. 155, 2011
697
Downloaded from on August 3, 2017 - Published by www.plantphysiol.org
Copyright © 2011 American Society of Plant Biologists. All rights reserved.
Timm et al.
Figure 6. HPR multiple mutants show reduced growth and total
chlorophyll content in ambient air. Plants were grown in ambient air
with a 10/14-h day/night cycle. Measurements were performed after
plants reached growth stadium 5.1 according to Boyes et al. (2001).
Rosette diameters were measured as the longest distance (in cm) across
fully expanded leaves. Mean values 6 SD are shown from five independent determinations each. Asterisks show significant changes
according to Student’s t test either in comparison with the corresponding wild type (* P , 0.05), with the single mutants (** P , 0.05), or with
the hpr1xhpr2 double mutant (*** P , 0.05). FW, Fresh weight.
more affected than the double mutant, with inhibition
levels up to 83% in comparison with the wild-type
control and a further growth retardation of up to 29%
compared with the double mutant. In clear contrast to
the growth reduction of the above-mentioned mutants, the single mutants hpr2 and hpr3 (compared with
the wild type) as well as the double knockouts
hpr1xhpr3 and hpr2xhpr3 (compared with hpr1 or hpr2
single mutant) displayed no growth effects.
Closely corresponding to the growth impairment,
total chlorophyll contents are reduced in HPR mutants
(Fig. 6). The hpr1 knockout has 35% less total leaf
chlorophyll content, while the hpr1xhpr2 double mutant contains 49% less total chlorophyll than the wild
type. This value is also decreased yet further in the
hpr1xhpr2xhpr3 triple mutant but not significantly so in
comparison with the hpr1xhpr2 double mutant. Furthermore, the leaf chlorophyll a/b ratio is also increased in the hpr1xhpr2 double mutant and in the
triple mutant, indicating a senescence-mediated chlorophyll breakdown (Pruzinská et al., 2005) that is
related to the elevated oxygen sensitivity. However,
neither hpr2 and hpr3 single mutant nor the corresponding double mutant hpr2xhpr3 are affected in
their chlorophyll contents or chlorophyll a/b ratios.
The third double mutant, hpr1xhpr3, only shows reduced levels in chlorophyll content comparable to hpr1
single mutants. These results once more clearly reflect
that HPR1 and HPR2 are the most active isoforms in
Arabidopsis.
As illustrated in Figure 5 (A and D, left panel), hpr1
single mutants also displayed a growth inhibition
phenotype under conditions in which the photoperiod
is shorter than 12 h of light. This effect is an exclusive
characteristic of hpr1 mutants, with hpr2 and hpr3
mutants not displaying such a photoperiodic effect on
growth (Fig. 5E, right panel). By contrast, the effect in
hpr1 single mutants vanishes upon light periods longer than 12 h of light (Fig. 5E, left panel; Timm et al.,
2008). This result possibly could be explained by
carbon starvation of the Calvin cycle under shorter
photoperiods during the prolonged night (Matt et al.,
1998). The photorespiratory flux is clearly not optimal
in hpr1 mutants under these conditions, which enhances the general short-day effect described above
and as suggested by decreased Glc levels at the end of
the night under short-day conditions (Fig. 7). Moreover, this effect could be reversed by cultivation in
high-carbon conditions (greater than 0.15% CO2),
clearly indicating that a disrupted photorespiratory
cycle is responsible for this daylength-dependent phenotype (Fig. 5D, right panel).
Photochemical Efficiencies Differ According to the
Number of Deleted HPRs
It is well known that repair mechanisms of PSII are
inhibited as a result of an interruption of the Calvin
cycle (Long et al., 1994; Takahashi and Murata, 2005).
To further substantiate the effects of HPR mutations on
PSII activity and hence limitations to the Calvin cycle,
we measured the photochemical efficiency (maximum
variable fluorescence/maximum yield of fluorescence
[Fv/Fm]) of PSII after different periods of exposure of
698
Plant Physiol. Vol. 155, 2011
Downloaded from on August 3, 2017 - Published by www.plantphysiol.org
Copyright © 2011 American Society of Plant Biologists. All rights reserved.
Photorespiratory Hydroxypyruvate Metabolism in Arabidopsis
Photosynthetic Parameters Differ According to the
Number of Deleted HPRs
Figure 7. HPR1 mutants show a reduced Glc level at the end of the
night under shorter photoperiods. Plants were grown in normal air with
10/14-h, 12/12-h, and 16/8-h day/night cycles. Leaves of three individual plants at developmental stage 5.1 (Boyes et al., 2001) were
harvested in the middle (MoD) and the end (EoD) of the day and the
middle (MoN) and the end (EoN) of the night. Mutant-to-wild-type
ratios of mean relative metabolite contents 6 SD are shown (n = 3), with
the mean wild-type values arbitrarily set to 1.
the mutants to ambient air (Fig. 8). Whereas wild-type
plants did not show any response, slightly decreased
values were observed for the hpr1 single mutant and
distinctly stronger effects for the hpr1xhpr2 double
knockout and the hpr1xhpr2xhpr3 triple mutant. This
effect was exacerbated with the length of exposure to
ambient air and is further exaggerated in the triple
mutant. In contrast, neither the hpr2 and hpr3 single
mutants nor the corresponding hpr2xhpr3 double
knockout showed a response in this parameter. The
double mutant hpr1xhpr3 displayed values comparable to hpr1 single mutants. Compared with other
photorespiratory mutants, in which the Fv/Fm ratio
responds very rapidly after exposure to ambient air
(Takahashi et al., 2007b), the respective HPR mutants
showed some degree of long-term response to air
levels of oxygen. This might be due to the relatively
moderate metabolic alterations when compared with
the changes observed after knockout of other photorespiratory enzymes.
To further characterize the response of HPR mutants
to various oxygen levels, which directly affect photorespiratory flux, we determined photosynthetic rates
and CO2 compensation points (g) at three different
oxygen concentrations (Fig. 9; Supplemental Table S3).
The second parameter, g, depends on the balance
between photosynthetic CO2 uptake and respiratory
CO2 release, and in C3 plants, its values respond
linearly to changing oxygen concentrations (Farquhar
et al., 1980). Since most of the multiple knockouts
show a distinct growth inhibition in air and an early
onset of flowering, HPR mutants were first cultivated
under mild nonphotorespiratory conditions (0.15%
CO2) to synchronize growth and then adapted to
ambient air for 1 week. Relative to wild-type plants,
we observed distinct though not always significant
changes to both parameters of photosynthetic gas
exchange in the order hpr3 , hpr2 , hpr1. This finding
indicates inferior but yet recognizable roles of HPR2
and HPR3 for photorespiratory metabolism as long as
HPR1 is active. In the absence of HPR1, the additional
deletion of HPR2 results in a massive negative impact
on photosynthesis that is even more pronounced in the
hpr1xhpr2xhpr3 triple mutant. While such effects were
observed at all three oxygen concentrations, they were
most noticeable at the highest oxygen concentration.
Additionally, the calculation of the slope (g) from the
estimated/oxygen response lines also indicates an
ascending negative effect on photosynthesis, according to the number of deleted HPRs. The value of g
increased in the order wild type , hpr1 , hpr1xhpr2 ,
hpr1xhpr2xhpr3 (numbers are given in the legend to
Fig. 9) if plants were first grown in high-carbon
conditions and then acclimatized to air conditions for
1 week. Control experiments confirmed that neither
HP nor Ser increased, nor were g values altered in the
mutants relative to the wild type before the transfer to
normal air (data not shown).
Leaves Accumulate Ser and Related Metabolites
To substantiate knockout-related effects under photorespiratory and nonphotorespiratory conditions, we
examined the levels of leaf amino acids using plants
grown in ambient air and elevated CO2 (0.15%). The
most striking observation was the strong accumulation of the photorespiratory intermediate Ser (Fig. 10).
The effect is consistent and rises with the level of HPR
inhibition, reaching as high as 26 mmol g21 fresh
weight in the hpr1xhpr2xhpr3 triple mutant. However,
not only Ser but also the metabolically related amino
acids Thr and Trp accumulate, suggesting an upregulation of Ser-metabolizing processes. Additionally,
Gly, the direct precursor of Ser within the photorespiratory cycle, also accumulates, underlining altered Ser
metabolism; however, this effect is not increased with
the level of HPR inhibition. Given that nearly all of the
Plant Physiol. Vol. 155, 2011
699
Downloaded from on August 3, 2017 - Published by www.plantphysiol.org
Copyright © 2011 American Society of Plant Biologists. All rights reserved.
Timm et al.
Figure 8. HPR multiple mutants show decreased Fv /Fm values after transfer to ambient air. Plants were grown in elevated
CO2 (0.15%) with a 10/14-h day/night
cycle. After reaching growth stadium 5.1
(Boyes et al., 2001), plants were transferred to ambient air. Fv/Fm ratios were
measured before transfer and 1, 2, and 3
weeks after transfer to normal air. Mean
values 6 SD (n = 5; five areas of interest
each) are shown. Asterisks show significant alterations according to Student’s t
test (P , 0.05 for the wild type [*], hpr1-1
[**], hpr1xhpr2 [***], and the corresponding line 1 week before transfer [+]).
described metabolic effects could be reversed by elevated CO2, it seems reasonable to assume that they
result from the disrupted photorespiratory flux.
DISCUSSION
During the last years, the core enzymes of the
photorespiratory cycle have been placed under intensive analyses at both the molecular and biochemical
levels (for review, see Bauwe et al., 2010). With the
exception of the multigene-encoded glycolate oxidase,
Arabidopsis mutants could be isolated for all of these
enzymes. Taken together, such mutants share one
common feature, their strong sensitivity to air, which
can easily be reverted by cultivation in elevated concentrations of CO2. This situation is long known from
studies with chemically generated mutants (Somerville, 2001), but, more recently, it was shown that
several exceptions exist. First, high-CO2 conditions
cannot rescue mutants without Gly decarboxylase
activity (Engel et al., 2007). This is because this particular enzyme not only functions in the photorespiratory cycle but also in one-carbon metabolism and
therefore is absolutely necessary for plant development under any conditions. Second, we previously
showed that, contrary to what was commonly thought,
photorespiratory HP reduction is not restricted to
peroxisomes but, via HPR2, can also occur in the
cytosol, resulting in an atypically low air sensitivity of
hpr1 mutants (Timm et al., 2008). Third, photorespiration has specific links to folate (Collakova et al.,
2008) and energy metabolism (Sweetlove et al., 2006).
The photorespiratory pathway is thus integrated into
whole-cell metabolism in a more complex manner
than previously considered. While most of the other
photorespiratory mutants are not very useful to investigate connections with other pathways, on account
of their conditional lethal phenotypes in ambient air,
the HPR mutants are not constrained in this manner.
Utilizing the moderate phenotype of HPR mutants,
connections to other pathways could be studied and
compared with the wild-type situation. Using this approach, we observed that knocking out HPR1 and
HPR2 invokes effects that are not restricted to photorespiration. In this previous study, we observed links
to tricarboxylic acid cycle intermediates, amino acid
metabolism, and ethanolamine, which acts as a precursor of choline biosynthesis (Timm et al., 2008). Such
cross talk and connections are still under investigation,
but the HPR mutants appear to be a very useful tool to
address such interactions.
This study presents evidence that HP reduction is
not even restricted to peroxisomes and cytosol but
involves chloroplasts too, albeit to a lesser extent. By
the identification of a novel 2-hydroxyacid dehydrogenase as an enzyme with HP and Glx reductase
activity (Fig. 1), we could enlarge the general importance of the reduction of these two metabolites. Since
the newly identified HPR3 is predicted to be located in
chloroplasts (http://aramemnon.botanik.uni-koeln.
de/tm_sub.ep?GeneID=13660&modelID=0; Schwacke
et al., 2003; Yu et al., 2008), a third compartment
is likely involved in the reduction of these photorespiratory intermediates. As we show by metabolite
analyses of HPR3 single knockouts (Fig. 3), some
photorespiratory intermediates (glycolate, Gly, Ser,
3HP, and glycerate) are slightly altered, which indicates a disruption of the photorespiratory cycle, thus
further cementing the association of HPR3 function to
700
Plant Physiol. Vol. 155, 2011
Downloaded from on August 3, 2017 - Published by www.plantphysiol.org
Copyright © 2011 American Society of Plant Biologists. All rights reserved.
Photorespiratory Hydroxypyruvate Metabolism in Arabidopsis
triple mutant, not only the CO2 compensation point
and net CO2 uptake were strongly inversely influenced (Fig. 9) but also the photochemical efficiency of
PSII after transfer of the mutant from elevated to air
levels of CO2 (Fig. 8).
This latter observation corresponds well to a previous report that impairment of the photorespiratory
cycle and, as a consequence, the Calvin cycle leads to
pronounced photoinhibition by suppression of the
synthesis of the D1 subunit of PSII (Takahashi et al.,
2007a). Apart from this impact on linear electron
transport, elevated levels of photorespiratory metabolites can inhibit carbon flux through the Calvin cycle.
Figure 9. Oxygen-dependent gas exchange of selected HPR mutants.
Plants were grown in elevated CO2 (0.15%) with a 10/14-h day/night
cycle. After reaching growth stadium 5.1 (Boyes et al., 2001), plants
were transferred to ambient air and, after 1 week of adaptation,
photosynthetic rates and CO2 compensation points were measured as
a function of oxygen concentration (10%, 21%, and 40%). Mean
values 6 SD (n = 5) are shown and are listed in Supplemental Table S3.
Asterisks show significant alterations according to Student’s t test (P ,
0.05 for the wild type [*], hpr3-1 [**], hpr2-1 [***], hpr1-1 [****], and
hpr1xhpr2 [*****]). Calculated slopes of the g/oxygen response lines
were 0.000355 6 0.000032 mL mL21 (wild type), 0.000507 6
0.000028 mL mL21 (hpr1), 0.000531 6 0.000024 mL mL21 (hpr1xhpr2),
and 0.000561 6 0.000018 mL mL21 (hpr1xhpr2xhpr3).
that of photorespiration. Furthermore, we discovered
changes in the levels of a-ketoglutarate, Glu, and Gln,
which indicate a shift in the carbon-nitrogen balance of
the mutants. These effects are consistent with those
observed following the knockout of other photorespiratory genes (Igarashi et al., 2006; Eisenhut et al.,
2008; Timm et al., 2008). The higher levels of Ala, Asn,
and Asp additionally indicate an impairment in nitrogen assimilation and may be regulated by the increased levels of Gln. The altered Thr level could
further support this notion or alternatively could be
due to the conversion of the Ser, which accumulates, to
alternative metabolites. Moreover, the combined deletion of HPR1, HPR2, and HPR3 leads to strong air
sensitivity, which, as mentioned above, is a cardinal
property of all other photorespiratory mutants. In this
Figure 10. Leaf amino acid analysis reveals strong accumulation of Ser
and related metabolites. Plants were grown in elevated CO2 (0.15%)
and in normal air with a 10/14-h day/night cycle. Leaves from six
individual plants at developmental stage 5.1 (Boyes et al., 2001) were
harvested in the middle of the photoperiod, and soluble amino acids
were analyzed by HPLC. Mean amino acid contents 6 SD are shown.
Asterisks indicate significant changes according to Student’s t test (P ,
0.05 for the corresponding wild-type sample [*], hpr1-1 [**], and
hpr1xhpr2 [***]). FW, Fresh weight.
Plant Physiol. Vol. 155, 2011
701
Downloaded from on August 3, 2017 - Published by www.plantphysiol.org
Copyright © 2011 American Society of Plant Biologists. All rights reserved.
Timm et al.
This is not restricted to the 2PG effects mentioned in
the introduction. It is also known, for example, that
Glx inhibits Rubisco activase and hence Rubisco
(Chastain and Ogren, 1989; Campbell and Ogren,
1990; Häusler et al., 1996), and Gly accumulation is
toxic due to binding of Mg2+ by complex formation
(Eisenhut et al., 2007). In addition, the elevated oxygen
dependency of g in all HPR mutants indicates higher
rates of decarboxylation of photorespiratory metabolites, for example Ser (Rontein et al., 2003) or HP
(Hedrick and Sallach, 1961), outside the core photorespiratory cycle. This is because the slope g of the
g-versus-oxygen response curve is essentially codetermined by two factors, the Rubisco specificity factor,
which is invariable in a given species, and the stoichiometry (i.e. tightness) of the photorespiratory cycle
(Farquhar et al., 1980). This stoichiometry of ribulose
1,5-bisphosphate oxygenation versus photorespiratory
CO2 release, which would be exactly 2 if only Gly
becomes decarboxylated, apparently shows some variation dependent on the fraction of additional decarboxylation reactions. This fraction could well be
enhanced by environmental conditions such as higher
light intensities and/or temperatures (Hanson and
Peterson, 1986) and the intactness of photorespiratory
metabolism (this work). Interestingly, mutants of the
peroxisomal malate dehydrogenases support these
results, because oxygen-dependent g values are also
changed in these plants (Cousins et al., 2008). Concerning the fact that both isoforms are responsible for
NADH supply of the peroxisomes, the double knockout is quite similar to HPR1 mutants and seems to be a
good confirmation of the results presented here. Unfortunately, no other defined photorespiratory mutants of Arabidopsis have been analyzed in this
context, and it remains open if this is an exclusive
feature of mutants with a defective HP-to-glycerate
conversion step.
The above-mentioned conclusion is further substantiated by the analysis of the leaf amino acid content of
the mutants, which revealed a distinct accumulation of
the photorespiratory intermediate Ser (Fig. 10). This is
perhaps not overly surprising, since this amino acid is
the direct precursor of HP (Liepman and Olsen, 2001),
but the enhanced levels could indeed trigger additional decarboxylating reactions, most prominently to
ethanolamine, as we have shown previously (Timm
et al., 2008), and therefore affect g in HPR mutants.
Nevertheless, this enhancement is not restricted only
to Ser but also includes the metabolically related
amino acids Thr and Trp and additionally influences
the steady-state level of Gly. The latter is consistently
increased in all mutants but showed no response to the
level of HPR mutation.
In addition to increased decarboxylating reactions of
photorespiratory metabolites, the ratio of respiration
in the light to the maximum rate of Rubisco carboxylation could also influence g. This also seems to be a
likely explanation for the altered values in HPR mutants. This notion is supported by a previously radio-
gasometric study of HPR1 mutants, which show
higher respiratory rates of primary and stored photosynthates in the dark and the light (supplemental table
S1 in Timm et al., 2008; Pärnik and Keerberg, 2007). It
seems reasonable that these reactions become pronounced with the level of HPR mutation and therefore
lead to increased values of g.
As we previously demonstrated, knocking out
HPR1 does not lead to the typical phenotype associated with the knockout of most other photorespiratory
enzymes (Timm et al., 2008). Interestingly, given the
daylength-dependent phenotype we demonstrated
here (Figs. 5 and 7), it seems likely that gene redundancy is only part of the reason behind this phenomenon. It is a general observation that plants that are
grown in short days become carbon starved during a
prolonged night (Matt et al., 1998; Graf et al., 2010).
This effect is exacerbated in HPR1 mutants, also indicating that there is a possible connection of photorespiration and carbon fixation during the Calvin cycle.
Most likely, this is mediated by reduced 3PGA recycling resulting from the disrupted photorespiratory
flux in combination with an outflow of metabolites
from the cycle. The photorespiratory flux is clearly not
optimal in hpr1 mutants under these conditions, which
enhances the general short-day effect described above
and as suggested by decreased Glc levels at the end of
the night under short-day conditions (Fig. 7).
Here, we report to our knowledge a previously
unknown enzyme that is able to reduce both HP and
Glx, two of the key metabolites within the photorespiratory cycle, and hence could directly or indirectly contribute to photorespiration in higher plants.
Considering its dual substrate specificity and its preference for NADPH as the cosubstrate, the identified
HPR3 could support peroxisomal and cytosolic HPRs
(Timm et al., 2008) as well as chloroplastidial and
cytosolic glyoxylate reductase (Allan et al., 2009) for an
optimal reduction of these intermediates. Given the
reported chloroplastic localization of HPR3 (Yu et al.,
2008), its exact physiological role is not absolutely
clear; however, the observed effects on photosynthesis
(Fig. 9) and the normalization of the elevated levels of
several diagnostic amino acids by elevated CO2 (Fig.
10) clearly suggest that its function is indeed associated with photorespiration. Unfortunately, no HP
transporter was identified yet (Reumann and Weber,
2006); therefore, it is not known how this photorespiratory intermediate gets into chloroplast. Since the
chemical structure of HP is similar to glycerate, it can be
hypothesized that import could possibly occur via the
reported glycerate/glycolate transporter (Howitz and
McCarty, 1986). In addition to the reduction of HP, the
newly identified HPR3 could also contribute to the
reduction of chloroplastidially generated Glx, which,
as discussed above, is an inhibitor of several enzymes
involved in photosynthetic CO2 fixation. This, however,
is not very likely, since a plastidial Glx reductase exists
(Simpson et al., 2008) and the single knockout of HPR3
does not cause drastic effects (Fig. 5).
702
Plant Physiol. Vol. 155, 2011
Downloaded from on August 3, 2017 - Published by www.plantphysiol.org
Copyright © 2011 American Society of Plant Biologists. All rights reserved.
Photorespiratory Hydroxypyruvate Metabolism in Arabidopsis
The construction of a chloroplast protein interaction
network in Arabidopsis emphasizes an additional role
of HPR3 in plant metabolism (Yu et al., 2008). From
these analyses, it appears that HPR3 could interact
with three other chloroplast proteins, two of which,
3PGA dehydrogenase (At1g17745) and phosphoserine
aminotransferase (At4g35630), are related to Ser synthesis. The third interacting protein, Ser hydroxymethyltransferase (At4g32520), is related to Gly synthesis
and one-carbon metabolism (Zhang et al., 2010). Thus,
HP could be generated in glycolytic routes from 3PGA,
most likely the phosphoserine pathway. This pathway
is best described in bacteria but also is found in plants,
mainly in heterotrophic issue or during the night when
the photorespiratory flux is not active (Ho et al., 1998;
Muñoz-Bertomeu et al., 2009). Supportive evidence for
this hypothesis is the accumulation of Gly, which we
observed in the HPR3 single mutants. If HPR3 works
in the phosphoserine pathway, the deletion of this
enzyme would be expected to lead to a disrupted flow,
which would explain the altered Gly-to-Ser interconversion in the mutants. However, the exact mechanism
behind these changes is not yet known, and it would
be very informative to analyze mutants within the
phosphoserine pathway under nonphotorespiratory
and photorespiratory conditions as well as in combination with different HPR mutations. It seems likely
that alternative Ser-synthesizing pathways, like the
phosphoserine-metabolizing route, interact with the
photorespiratory pathway and, therefore, the HPR
triple mutant shows further growth retardation.
In conclusion, although the role for HPR3 will need
further investigation, this study presents strong
indications that the enzyme is associated with the
photorespiratory process and can at least partially
complement the function of the peroxisomal HPR1
and the cytosolic HPR2. Moreover, HPR3 could display a possible link of photorespiration to other Sersynthesizing pathways.
MATERIALS AND METHODS
Plant Material and Growth
Arabidopsis (Arabidopsis thaliana) ecotypes Columbia and Landsberg erecta
were used as wild-type references. SALK lines SALK_067724 (hpr1-1),
SALK_143584 (hpr1-2), SALK_105875 (hpr2-1), SALK_019014 (hpr3-1), and
SALK_143689 (hpr3-2) and gene trap line GT.89166 (hpr2-2) were obtained
from the Nottingham Arabidopsis Stock Centre (Sundaresan et al., 1995;
Alonso et al., 2003). Seeds were incubated at 4°C for at least 2 d to break
dormancy prior to germination. Plants were grown on soil (Type Mini Tray;
Einheitserdewerk) and vermiculite (4:1 mixture) and were regularly watered
with 0.2% Wuxal liquid fertilizer (Aglukon). Unless otherwise stated, plants
were grown in normal air (380–400 mL L21 CO2) and in air with elevated CO2
(0.15% or 1% as indicated in the figure legends) in controlled-environment
chambers (Percival) with a day/night cycle of 10/14 h (22°C/18°C, approximately 120 mmol m22 s21 irradiance).
Knockout Mutants and RT-PCR
Insertion lines for HPR1 (hpr1-1 and hpr1-2) and HPR2 (hpr2-1 and hpr2-2)
were described previously (Timm et al., 2008). For T-DNA insertion lines of
HPR3, leaf DNA was isolated and PCR amplified (1 min at 94°C, 1 min at 58°C,
1 min at 72°C; 35 cycles) with specific left border primers (R295 for hpr3-1 and
R175 for hpr3-2) and a corresponding primer for the open reading frame of
At1g12550 (R669) to verify insertions. The nucleotide sequences of all primers
used in this study are shown in Supplemental Table S4. Zygosity was then
validated using leaf DNA and specific primers for the genomic region of
At1g12550 (R665 and R669). The knockout of HPR3 was verified by RT-PCR
using 2.5 mg of leaf RNA for cDNA synthesis (Nucleospin RNA plant kit
[Macherey-Nagel] and RevertAid cDNA synthesis kit [MBI Fermentas]) and
the primer combination P424/P425, yielding a 972-bp fragment. This fragment was present in the wild type but not in homozygous knockout lines. The
amount of cDNA for PCR was calibrated according to signals from 432-bp
fragments obtained by PCR amplification of the constitutively expressed 40S
ribosomal protein S16 gene, with primers R176 (sense) and R177 (antisense)
before PCR analysis.
Multiple HPR mutants were generated by crossing the HPR1 and HPR2
double knockout (Timm et al., 2008) with the isolated line hpr3-1, yielding a
triple heterozygous intermediate line for HPR1, HPR2, and HPR3. This line
then was grown under elevated CO2 concentrations (1%). Leaf DNA was
isolated and analyzed by genomic PCR to verify the presence of all three
insertions and zygosity using the primers listed in Timm et al. (2008) for HPR1
and HPR2 and in Supplemental Table S4 for HPR3. The complete knockout of
HPRs was double checked by RT-PCR as described above, yielding two
double mutants, hpr1-1xhpr3-1 and hpr2-1xhpr3-1, and the corresponding
triple mutant hpr1-1xhpr2-1xhpr3-1.
Heterologous Expression
The entire coding region of At1g12550 transcript was amplified from leaf
RNA by RT-PCR with primers R987 and R988 (Supplemental Table S4) under
simultaneous introduction of two flanking SacII restriction sites. The PCR
product was first cloned into pGEM T-Easy (Promega) and then, after
sequence confirmation and using the flanking SacII restriction sites, inserted
in-frame into the coding sequence of the N-terminal His tag of the Escherichia
coli overexpression vector pHUE (Catanzariti et al., 2004). Recombinant
HPR3 was purified using nickel-nitrilotriacetic acid agarose (HisTrap) and
Q-Sepharose blue affinity chromatography (GE Healthcare).
Chlorophyll Determination and PSII Fluorescence
Leaf chlorophyll content was determined as described by Porra (2002) and
Porra et al. (1989). The Fv/Fm ratio corresponding to the potential quantum
yield of photochemical reactions of PSII was measured (Schreiber et al., 1986)
using an imaging pulse amplitude modulation fluorometer (M series; Walz).
Plants were first grown for 8 weeks under elevated CO2 conditions (1%) and
then transferred to ambient air (10/14-h day/night cycle). Measurements
were performed before (high-carbon control) and after transfer to ambient air
(after 1, 2, and 3 weeks) as described by Fahnenstich et al. (2008).
CO2 Exchange
Gas exchange (net photosynthetic rate and CO2 compensation point) was
analyzed with a Li-Cor-6400 gas exchange system (LI-COR) using fully
developed rosette leaves from plants grown in normal air or air with elevated
CO2 (0.15%) to developmental stage 5.1 as described above. Measurements
were performed at a photosynthetic photon flux density of 250 mmol m22 s21,
supplied by an in-built red/blue light-emitting diode light source, 380 mL L21
CO2, 21% O2, and a leaf temperature of 25°C. Alternative oxygen concentrations (10% and 40%) were generated using the gas-mixing system GMS600
(QCAL Messtechnik). Plants were adapted to the respective conditions for at
least 20 min before the actual measurement.
Metabolomics and Leaf Soluble Amino Acids
Metabolite analysis was performed using rosette leaves from plants at
developmental stage 5.1 according to Boyes et al. (2001). Extraction and
analysis were performed as described by Lisec et al. (2006). For amino acid
determination, 100 mg of leaf material was ground in liquid nitrogen and
extracted in 1.8 mL of 80% ethanol for 30 min. After centrifugation, the
supernatant was vacuum dried and dissolved in 8 mM sodium phosphate (pH
6.8) and 0.4% tetrahydrofurane. Individual amino acids were separated by
HPLC and quantified as described earlier (Hagemann et al., 2005).
Plant Physiol. Vol. 155, 2011
703
Downloaded from on August 3, 2017 - Published by www.plantphysiol.org
Copyright © 2011 American Society of Plant Biologists. All rights reserved.
Timm et al.
Sequence data from this article can be found in the Arabidopsis Genome
Initiative (2000) or GenBank/EMBL data libraries under the following accession numbers: At1g68010 (HPR1), At1g79870 (HPR2), At1g12550 (HPR3), and
At2g09990 (40S ribosomal protein S16).
Supplemental Data
The following materials are available in the online version of this article.
Supplemental Table S1. Relative levels of selected metabolites in the
single mutants of HPR3.
Supplemental Table S2. Redistribution of label following feeding of leaf
discs with [U-13C]Glc in the individual hpr3-1 and hpr3-2 knockout
plants.
Supplemental Table S3. Oxygen-dependent gas-exchange parameters of
HPR mutants.
Supplemental Table S4. Sequences of primers used in this study.
ACKNOWLEDGMENTS
We are especially grateful to Martin Hagemann (University of Rostock) for
helpful comments on the research. The excellent technical assistance with
HPLC analysis from Klaudia Michl (University of Rostock) is acknowledged.
We further thank the Nottingham Arabidopsis Stock Centre for providing us
with the SALK lines to make this work possible.
Received September 23, 2010; accepted December 21, 2010; published December 23, 2010.
LITERATURE CITED
Allan WL, Clark SM, Hoover GJ, Shelp BJ (2009) Role of plant glyoxylate
reductases during stress: a hypothesis. Biochem J 423: 15–22
Alonso JM, Stepanova AN, Leisse TJ, Kim CJ, Chen H, Shinn P, Stevenson
DK, Zimmerman J, Barajas P, Cheuk R, et al (2003) Genome-wide insertional mutagenesis of Arabidopsis thaliana. Science 301: 653–657
Anderson LE (1971) Chloroplast and cytoplasmic enzymes. II. Pea leaf
triose phosphate isomerases. Biochim Biophys Acta 235: 237–244
Arabidopsis Genome Initiative (2000) Analysis of the genome sequence of
the flowering plant Arabidopsis thaliana. Nature 408: 796–815
Bauwe H, Hagemann M, Fernie AR (2010) Photorespiration: players,
partners and origin. Trends Plant Sci 15: 330–336
Boldt R, Edner C, Kolukisaoglu Ü, Hagemann M, Weckwerth W, Wienkoop S,
Morgenthal K, Bauwe H (2005) D-GLYCERATE 3-KINASE, the last unknown enzyme in the photorespiratory cycle in Arabidopsis, belongs to a
novel kinase family. Plant Cell 17: 2413–2420
Boyes DC, Zayed AM, Ascenzi R, McCaskill AJ, Hoffman NE, Davis KR,
Görlach J (2001) Growth stage-based phenotypic analysis of Arabidopsis:
a model for high throughput functional genomics in plants. Plant Cell
13: 1499–1510
Campbell WJ, Ogren WL (1990) Glyoxylate inhibition of ribulosebisphosphate carboxylase/oxygenase activation in intact, lysed, and reconstituted chloroplasts. Photosynth Res 23: 257–268
Catanzariti AM, Soboleva TA, Jans DA, Board PG, Baker RT (2004) An
efficient system for high-level expression and easy purification of
authentic recombinant proteins. Protein Sci 13: 1331–1339
Chastain CJ, Ogren WL (1989) Glyoxylate inhibition of ribulosebisphosphate carboxylase/oxygenase activation state in vivo. Plant Cell Physiol
30: 937–944
Collakova E, Goyer A, Naponelli V, Krassovskaya I, Gregory JF III,
Hanson AD, Shachar-Hill Y (2008) Arabidopsis 10-formyl tetrahydrofolate deformylases are essential for photorespiration. Plant Cell 20:
1818–1832
Cousins AB, Pracharoenwattana I, Zhou W, Smith SM, Badger MR (2008)
Peroxisomal malate dehydrogenase is not essential for photorespiration
in Arabidopsis but its absence causes an increase in the stoichiometry of
photorespiratory CO2 release. Plant Physiol 148: 786–795
Eisenhut M, Bauwe H, Hagemann M (2007) Glycine accumulation is toxic
for the cyanobacterium Synechocystis sp. strain PCC 6803, but can be
compensated by supplementation with magnesium ions. FEMS Microbiol Lett 277: 232–237
Eisenhut M, Huege J, Schwarz D, Bauwe H, Kopka J, Hagemann M (2008)
Metabolome phenotyping of inorganic carbon limitation in cells of the
wild type and photorespiratory mutants of the cyanobacterium Synechocystis sp. strain PCC 6803. Plant Physiol 148: 2109–2120
Engel N, van den Daele K, Kolukisaoglu Ü, Morgenthal K, Weckwerth W,
Pärnik T, Keerberg O, Bauwe H (2007) Deletion of glycine decarboxylase in Arabidopsis is lethal under nonphotorespiratory conditions.
Plant Physiol 144: 1328–1335
Fahnenstich H, Scarpeci TE, Valle EM, Flugge UI, Maurino VG (2008)
Generation of hydrogen peroxide in chloroplasts of Arabidopsis overexpressing glycolate oxidase as an inducible system to study oxidative
stress. Plant Physiol 148: 719–729
Farquhar GD, von Caemmerer S, Berry JA (1980) A biochemical model
of photosynthetic CO2 assimilation in leaves of C3 species. Planta 149:
78–90
Graf A, Schlereth A, Stitt M, Smith AM (2010) Circadian control of
carbohydrate availability for growth in Arabidopsis plants at night.
Proc Natl Acad Sci USA 107: 9458–9463
Hagemann M, Vinnemeier J, Oberpichler I, Boldt R, Bauwe H (2005) The
glycine decarboxylase complex is not essential for the cyanobacterium
Synechocystis sp. strain PCC 6803. Plant Biol (Stuttg) 7: 15–22
Hanson KR, Peterson RB (1986) Regulation of photorespiration in leaves:
evidence that the fraction of ribulose bisphosphate oxygenated is
conserved and stoichiometry fluctuates. Arch Biochem Biophys 246:
332–346
Häusler RE, Bailey KJ, Lea PJ, Leegood RC (1996) Control of photosynthesis in barley mutants with reduced activities of glutamine synthetase
and glutamate synthase. 3. Aspects of glyoxylate metabolism and effects of glyoxylate on the activation state of ribulose-1,5-bisphosphate
carboxylase-oxygenase. Planta 200: 388–396
Hedrick JL, Sallach HJ (1961) The metabolism of hydroxypyruvate. I. The
nonenzymatic decarboxylation and autoxidation of hydroxypyruvate.
J Biol Chem 236: 1867–1871
Ho CL, Noji M, Saito M, Yamazaki M, Saito K (1998) Molecular characterization of plastidic phosphoserine aminotransferase in serine biosynthesis from Arabidopsis. Plant J 16: 443–452
Howitz KT, McCarty RE (1986) D-Glycerate transport by the pea chloroplast glycolate carrier: studies on [1-14C]D-glycerate uptake and
D-glycerate dependent O2 evolution. Plant Physiol 80: 390–395
Igarashi D, Miwa T, Seki M, Kobayashi M, Kato T, Tabata S, Shinozaki
K, Ohsumi C (2003) Identification of photorespiratory glutamate:
glyoxylate aminotransferase (GGAT) gene in Arabidopsis. Plant J 33:
975–987
Igarashi D, Tsuchida H, Miyao M, Ohsumi C (2006) Glutamate:glyoxylate
aminotransferase modulates amino acid content during photorespiration. Plant Physiol 142: 901–910
Kelly GJ, Latzko E (1976) Inhibition of spinach-leaf phosphofructokinase
by 2-phosphoglycollate. FEBS Lett 68: 55–58
Kleczkowski LA, Randall DD, Edwards GE (1991) Oxalate as a potent and
selective inhibitor of spinach (Spinacia oleracea) leaf NADPH-dependent
hydroxypyruvate reductase. Biochem J 276: 125–127
Liepman AH, Olsen LJ (2001) Peroxisomal alanine:glyoxylate aminotransferase (AGT1) is a photorespiratory enzyme with multiple substrates in
Arabidopsis thaliana. Plant J 25: 487–498
Lisec J, Schauer N, Kopka J, Willmitzer L, Fernie AR (2006) Gas chromatography mass spectrometry-based metabolite profiling in plants. Nat
Protoc 1: 387–396
Long SP, Humphries S, Falkowski PG (1994) Photoinhibition of photosynthesis in nature. Annu Rev Plant Physiol 45: 633–662
Matt P, Schurr U, Klein D, Krapp A, Stitt M (1998) Growth of tobacco in
short-day conditions leads to high starch, low sugars, altered diurnal
changes in the Nia transcript and low nitrate reductase activity, and
inhibition of amino acid synthesis. Planta 207: 27–41
Muñoz-Bertomeu J, Cascales-Miñana B, Mulet JM, Baroja-Fernández E,
Pozueta-Romero J, Kuhn JM, Segura J, Ros R (2009) Plastidial glyceraldehyde-3-phosphate dehydrogenase deficiency leads to altered root
development and affects the sugar and amino acid balance in Arabidopsis. Plant Physiol 151: 541–558
Pärnik T, Keerberg O (2007) Advanced radiogasometric method for the
determination of the rates of photorespiratory and respiratory decar-
704
Plant Physiol. Vol. 155, 2011
Downloaded from on August 3, 2017 - Published by www.plantphysiol.org
Copyright © 2011 American Society of Plant Biologists. All rights reserved.
Photorespiratory Hydroxypyruvate Metabolism in Arabidopsis
boxylations of primary and stored photosynthates under steady-state
photosynthesis. Physiol Plant 129: 34–44
Porra RJ (2002) The chequered history of the development and use of
simultaneous equations for the accurate determination of chlorophylls a
and b. Photosynth Res 73: 149–156
Porra RJ, Thompson WA, Kriedemann PE (1989) Determination of accurate extinction coefficients and simultaneous equations for assaying
chlorophyll a and chlorophyll b extracted with 4 different solvents:
verification of the concentration of chlorophyll standards by atomic
absorption spectroscopy. Biochim Biophys Acta 975: 384–394
Pruzinská A, Tanner G, Aubry S, Anders I, Moser S, Müller T, Ongania
KH, Kräutler B, Youn JY, Liljegren SJ, et al (2005) Chlorophyll breakdown in senescent Arabidopsis leaves: characterization of chlorophyll
catabolites and of chlorophyll catabolic enzymes involved in the
degreening reaction. Plant Physiol 139: 52–63
Reumann S, Weber AP (2006) Plant peroxisomes respire in the light: some
gaps of the photorespiratory C2 cycle have become filled—others remain. Biochim Biophys Acta 1763: 1496–1510
Rontein D, Rhodes D, Hanson AD (2003) Evidence from engineering that
decarboxylation of free serine is the major source of ethanolamine
moieties in plants. Plant Cell Physiol 44: 1185–1191
Schreiber U, Schliwa U, Bilger W (1986) Continuous recording of photochemical and nonphotochemical chlorophyll fluorescence quenching
with a new type of modulation fluorometer. Photosynth Res 10: 51–62
Schwacke R, Schneider A, van der Graaff E, Fischer K, Catoni E,
Desimone M, Frommer WB, Flügge UI, Kunze R (2003) ARAMEMNON, a novel database for Arabidopsis integral membrane proteins.
Plant Physiol 131: 16–26
Schwarte S, Bauwe H (2007) Identification of the photorespiratory
2-phosphoglycolate phosphatase, PGLP1, in Arabidopsis. Plant Physiol
144: 1580–1586
Sienkiewicz-Porzucek A, Nunes-Nesi A, Sulpice R, Lisec J, Centeno DC,
Carillo P, Leisse A, Urbanczyk-Wochniak E, Fernie AR (2008) Mild
reductions in mitochondrial citrate synthase activity result in a compromised nitrate assimilation and reduced leaf pigmentation but have
no effect on photosynthetic performance or growth. Plant Physiol 147:
115–127
Simpson JP, Di Leo R, Dhanoa PK, Allan WL, Makhmoudova A, Clark
SM, Hoover GJ, Mullen RT, Shelp BJ (2008) Identification and characterization of a plastid-localized Arabidopsis glyoxylate reductase isoform: comparison with a cytosolic isoform and implications for cellular
redox homeostasis and aldehyde detoxification. J Exp Bot 59: 2545–2554
Somerville CR (2001) An early Arabidopsis demonstration: resolving a few
issues concerning photorespiration. Plant Physiol 125: 20–24
Sundaresan V, Springer P, Volpe T, Haward S, Jones JD, Dean C, Ma H,
Martienssen R (1995) Patterns of gene action in plant development
revealed by enhancer trap and gene trap transposable elements. Genes
Dev 9: 1797–1810
Sweetlove LJ, Lytovchenko A, Morgan M, Nunes-Nesi A, Taylor NL,
Baxter CJ, Eickmeier I, Fernie AR (2006) Mitochondrial uncoupling
protein is required for efficient photosynthesis. Proc Natl Acad Sci USA
103: 19587–19592
Takahashi S, Bauwe H, Badger M (2007a) The photorespiratory pathway is
important for the repair of photodamaged photosystem II under CO2
limiting conditions. Photosynth Res 91: 209
Takahashi S, Bauwe H, Badger M (2007b) Impairment of the photorespiratory pathway accelerates photoinhibition of photosystem II by suppression of repair but not acceleration of damage processes in
Arabidopsis. Plant Physiol 144: 487–494
Takahashi S, Murata N (2005) Interruption of the Calvin cycle inhibits the
repair of photosystem II from photodamage. Biochim Biophys Acta
1708: 352–361
Timm S, Nunes-Nesi A, Pärnik T, Morgenthal K, Wienkoop S, Keerberg
O, Weckwerth W, Kleczkowski LA, Fernie AR, Bauwe H (2008) A
cytosolic pathway for the conversion of hydroxypyruvate to glycerate
during photorespiration in Arabidopsis. Plant Cell 20: 2848–2859
Voll LM, Jamai A, Renné P, Voll H, McClung CR, Weber APM (2006) The
photorespiratory Arabidopsis shm1 mutant is deficient in SHM1. Plant
Physiol 140: 59–66
Yu QB, Li G, Wang G, Sun JC, Wang PC, Wang C, Mi HL, Ma WM, Cui J,
Cui YL, et al (2008) Construction of a chloroplast protein interaction
network and functional mining of photosynthetic proteins in Arabidopsis
thaliana. Cell Res 18: 1007–1019
Zhang Y, Sun K, Sandoval FJ, Santiago K, Roje S (2010) One-carbon
metabolism in plants: characterization of a plastid serine hydroxymethyltransferase. Biochem J 430: 97–105
Plant Physiol. Vol. 155, 2011
705
Downloaded from on August 3, 2017 - Published by www.plantphysiol.org
Copyright © 2011 American Society of Plant Biologists. All rights reserved.